Next Article in Journal
Systematic Understanding of Mechanisms of a Chinese Herbal Formula in Treatment of Metabolic Syndrome by an Integrated Pharmacology Approach
Next Article in Special Issue
High-Throughput Screening Methodology to Identify Alpha-Synuclein Aggregation Inhibitors
Previous Article in Journal
Characterization of Starch Degradation Related Genes in Postharvest Kiwifruit
Previous Article in Special Issue
Peculiarities of the Super-Folder GFP Folding in a Crowded Milieu
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress in Machine Learning-Based Methods for Protein Fold Recognition

School of Computer Science and Technology, Tianjin University, Tianjin 300354, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2016, 17(12), 2118; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17122118
Submission received: 16 November 2016 / Revised: 3 December 2016 / Accepted: 11 December 2016 / Published: 16 December 2016
(This article belongs to the Special Issue Protein Folding)

Abstract

:
Knowledge on protein folding has a profound impact on understanding the heterogeneity and molecular function of proteins, further facilitating drug design. Predicting the 3D structure (fold) of a protein is a key problem in molecular biology. Determination of the fold of a protein mainly relies on molecular experimental methods. With the development of next-generation sequencing techniques, the discovery of new protein sequences has been rapidly increasing. With such a great number of proteins, the use of experimental techniques to determine protein folding is extremely difficult because these techniques are time consuming and expensive. Thus, developing computational prediction methods that can automatically, rapidly, and accurately classify unknown protein sequences into specific fold categories is urgently needed. Computational recognition of protein folds has been a recent research hotspot in bioinformatics and computational biology. Many computational efforts have been made, generating a variety of computational prediction methods. In this review, we conduct a comprehensive survey of recent computational methods, especially machine learning-based methods, for protein fold recognition. This review is anticipated to assist researchers in their pursuit to systematically understand the computational recognition of protein folds.

1. Introduction

Understanding how proteins adopt their 3D structure remains one of the greatest challenges in science. Elucidation of this process would greatly impact various fields of biology and medicine, as well as the rational design of new functional proteins and drug molecules. Determination of the fold category of a protein is crucial as it reveals the 3D structure of proteins. Classification of a protein of unknown structure under a fold category is called fold recognition, which is a fundamental step in the determination of the tertiary structure of a protein.
In the early years, determination of protein structure relies on traditional experimental methods, such as X-ray crystallography and nuclear magnetic resonance spectroscopy. In the post-genomic era, numerous sequences are generated by next-generation sequencing techniques. Although an increasing number of sequences are structurally characterized using experimental methods, the gap between structurally determined sequences and uncharacterized sequences is constantly increasing. Therefore, developing computational methods for fast and accurate determination of protein structures is urgently needed. Accurate computational prediction of protein folds has recently emerged as alternative approach to the labor intensive and expensive experimental methods. Computational methods for protein fold recognition can be generally categorized into three classes: (1) de novo modeling methods; (2) template-based methods; and (3) template-free methods. Many efforts have focused on the development of methods under classes (2) and (3) because the de novo approach (class 1) has two limitations. First, it requires long computational time and numerous sources, and second, it can only be successfully applied in small proteins.
Template-based methods used to determine protein structures are based on the evolutionary relationships of proteins. The procedure for template-based methods can be summarized as follows: First, proteins of known structures retrieved from public protein structure databases (e.g., Protein Data Bank (PDB)) are used as template proteins for a query protein sequence. To make template-based prediction fast and reliable, a simplified database is usually employed, in which the sequence similarity is less than 50%–70%. Second, distant evolutionary relationships between a target sequence and proteins of known structure are detected. In this step, multi-alignment algorithms are adopted to exploit evolutionary information by encoding amino acid sequences into profiles. Third, to determine the optimal alignments, scoring functions are usually used as measures to evaluate the similarity between the profiles derived from a query protein and those of template proteins with known structures. Z-score and E-value are the two commonly used scoring functions. The accuracy of the alignment is tremendously important in model building. Fourth, 3D structure models based on template atom coordinates and optimal query-template alignments are built. Last, the optimal structure models are determined from the model candidates through further structure optimization. The commonly used structural optimization methods include energy minimization and loop modeling.
A series of template-based methods were developed in the last few decades. This series of approaches are regarded as the most successful methods in constructing theoretical models of protein structures. For instance, Jaroszewski et al. [1] developed a protein recognition method called Fold and Function Assignment System (FFAS) by using a profile-profile alignment strategy without using any structural information. In FFAS, query and template profiles are obtained by PSI-BLAST searching against the NR85 database; these profiles are then aligned by a dot-product scoring function. The significance of alignment scores was calculated by comparing the protein with the distribution scores from pairs of unrelated proteins. Xu et al. [2] improved the FFAS method and proposed a method called FFAS-3D, wherein they introduced structural information, such as secondary structure, solvent accessibility, and residue depth. FFAS-3D remarkably outperforms FFAS. Moreover, Shi et al. [3] developed a protein fold recognition method called FUGUE, which can search sequences against protein fold libraries by using environment-specific substitution tables and structure-dependent gap penalties. Raptor is a novel method that uses the mathematical theory of linear programming to build 3D models of proteins and predict protein folds [4,5]. Roy et al. [6] developed an online prediction server called I-TASSER (Iterative Threading ASSEmbly Refinement), which is an integrated platform for automated protein structure and function prediction based on the sequence-to-structure-to-function paradigm. Ghouzam et al. [7] proposed ORION, a new fold recognition method based on pairwise comparison of hybrid profiles that contain evolutionary information from protein sequences and their structures. Other template-based methods were successfully developed, including MODELLER [8] and TMFR [9]. MODELLER implements comparative protein structure modeling through satisfaction of spatial restraints, whereas TMFR applies special scoring functions to align sequences and predict whether given sequence pairs share the same fold. As mentioned above, several typical template-based methods have been proposed. However, the manner by which to examine the quality of template-based modeling methods remains unknown. Currently, CASP (Critical Assessment of protein Structure Prediction) is a mainstream platform used to establish an independent mechanism to assess the current methods employed in protein structure modeling [10]. This platform can be accessed at http://predictioncenter.org/.
Although much progress has been made in template-based methods, some problems still exist, as follows: First, we need to determine the structures of template proteins. The three-dimensional structures of many proteins remain to be determined. Second, template-based modelling largely relies on the homology between target and template proteins. When the target and template proteins display a sequence similarity of >30%, the use of sequence alignment methods (e.g., BLAST [11] and SSEARCH [12]) can reveal their evolutionary relationships. However, this approach is not available for non-obvious relationships between targets and templates with a sequence identity of lower than 20%–30%. Third, template-based structure modeling is time consuming. This approach always requires homology detection by searching target proteins against a template database to detect distant evolutionary relationships.
To address the aforementioned problems, recent research efforts have focused on the development of template-free methods. Template-free methods seek to build models and accurately predict protein structures solely based on amino acid sequences rather than on known structural proteins as templates. Many machine learning algorithms have been recently used for that purpose; these algorithms include Hidden Markov Model (HMM), genetic algorithm, Artificial Neural Network, Support Vector Machines (SVMs), and ensemble classifiers. A key underlying assumption in employing machine learning-based methods for protein fold recognition is that the number of protein fold classes is limited [13]. Machine learning aims to build a prediction model by learning the differences between different protein fold categories and use the learned model to automatically assign a query protein to a specific protein fold class. This approach is thus more efficient for large-scale predictions and can examine a large number of promising candidates for further experimental validation. This review focuses mainly on the recent progress in machine learning-based methods for protein fold recognition. This review is organized as follows: First, we introduce the public databases usually used in protein fold recognition research. Second, we describe the framework and flowchart of machine learning-based recognition methods. Third, we summarize some recent representative machine learning-based methods for protein fold recognition. Finally, we evaluate and compare the recognition performance of existing methods used in the last 10 years on a benchmark dataset.

2. Databases

Multiple database sources are often used in protein structure research. These databases include PDB [14]; Universal Protein Resource [15]; Database of Secondary Structure of Protein (DSSP) [16]; Structural Classification of Proteins (SCOP) [17]; SCOP2 (a successor of SCOP) [18]; and Class, Architecture, Topology, Homology (CATH) [19] (Table 1). Among these databases, SCOP and CATH have become valuable resources in protein fold recognition research. Figure 1 shows the architectures of these databases. These databases are detailed below.

2.1. SCOP and SCOP2

SCOP, proposed by Murzin et al. [14], is a hierarchical protein classification database that aims to organize structurally characterized proteins based on their structural and evolutionary relationships. Proteins in SCOP are categorized into four hierarchical levels: family, superfamily, protein fold, and structural class. At the family level, proteins are clustered into families based on one of two principles; the first principle is that proteins display more than 30% sequence identity, and the second is that the proteins with lower sequence identities share similar structure and functions. Families containing proteins with low sequence identities but with similar structural and functional features and sharing a common evolutionary origin are grouped into superfamilies. At the fold level, superfamilies and families are clustered into a fold if their proteins display the same secondary structures in the same arrangement with similar topological connections. At the structural class level, different folds are grouped into classes for the convenience of users. In SCOP, seven different structural classes are formed based on protein secondary structure contents: (1) all α; (2) all-β; (3) α and β; (4) α plus β; (5) multi-domain proteins; (6) membrane and cell surface proteins; and (7) small proteins.
Murzin et al. [14] have recently presented a successor of SCOP, called SCOP2, which is available at http://scop2.mrc-lmb.cam.ac.uk. Compared with SCOP, SCOP2 displays a more advanced framework for protein structure classification, wherein the best features of SCOP are retained and a novel approach for classification of protein structures is offered. In SCOP2, protein sequences and their structures are presented in a directed acyclic graph to form a network of many-to-many relationships.

2.2. CATH

Similar to SCOP, CATH is a hierarchical protein domain classification. In the CATH database, proteins and their structures are obtained from the PDB database. When proteins share a clear common evolutionary ancestor, they are clustered into a homologous superfamily (“H” level in CATH, Figure 1). When proteins in the same homologous superfamily display the same fold but do not obviously show evolutionary relationships, they are grouped into the same topology (“T” level). Proteins in the “T” level show similar secondary structural arrangements and are clustered into the same architecture (“A” level). For that end, the architectures are further grouped into structural classes (“C” level) according to secondary structure content.

3. Framework of Machine Learning-Based Methods

This section describes the mechanism of protein fold recognition by machine learning-based methods. The overall procedure in protein fold recognition by machine learning-based methods includes two phases (Figure 2): (1) model training; and (2) prediction.
In the first phase (model building), query protein sequences are first submitted into a pipeline of feature representation, in which sequences of different lengths are encoded with fixed-length feature vectors by feature descriptors. The commonly used feature descriptors include Amino Acid Composition (AAC), Pseudo AAC, Functional Domain (FunD), Position Specific Scoring Matrix (PSSM)-based descriptors, Secondary Structure-based descriptors, and Autocross-covariance (ACC) transformation. When the resulting feature representations display some irrelevant features or redundant features, an alternative step is usually performed to select the optimal feature subsets, which can yield the best performance, from the resulting feature representations. Subsequently, the feature vectors are fed into a pre-selected classification algorithm to train a prediction model. Typical classification algorithms often used in model building include SVM, Random Forest (RF), Naïve Bayes (NB), and Logistic Regression (LR). The first phase is completed in this step.
In the second phase (prediction), uncharacterized query proteins are first submitted into the same pipeline of feature representation as in the first phase. Note that if feature optimization of the generated feature representation is performed in the first phase, feature optimization should also be performed in the second phase; otherwise, the resulting feature vectors are fed into the trained prediction model, wherein the protein fold class to which the query proteins belong is predicted.

4. Recent Representative Methods for Protein Fold Recognition

Machine learning-based methods can be further categorized into two classes according to the learning algorithms used in prediction models: (1) single classifier-based methods; and (2) ensemble classifier-based methods. Single classifier-based methods use a single specific learning algorithm to build prediction models, whereas ensemble classifier-based methods use an ensemble of multiple, either similar or different, learning algorithms to build prediction models. This section introduces the recent single classifier-based methods and ensemble classifier-based methods used in protein fold recognition, as follows:

4.1. Single Classifier-Based Methods

Most of the current single classifier methods used in protein fold recognition are based on SVM classifier probably because SVM, being a well-known classification algorithm, has been highly efficient in several fields of bioinformatics, such as in protein remote homology detection, protein structural classification, and DNA-binding protein prediction. SVM-based protein fold recognition methods include Shamim’s method [20], Damoulas’ method [21], ACCFold_AC and ACCFold_ACC [22], TAXFOLD [23], and Alok Sharma’s method [24]. The main difference among these SVM-based methods is their feature representation algorithms. For instance, Shamim et al. [20] propose to use secondary structural state and solvent accessibility state frequencies of amino acids and amino acid pairs as feature vectors. Of these features, the secondary structural state frequencies are more effective than the two other features for fold class discrimination. Combining the secondary structural state frequencies with the two other feature types can further improve the accuracy of fold discrimination. Dong et al. [22] propose the ACCFold_AC and ACCFold_ACC methods for protein fold recognition. Based on the distant evolutionary relationships of protein sequences, their proposed feature algorithm can effectively capture the evolutionary information embedded in the form of position-specific score matrices and the sequence-order effect by utilizing ACC transformation. The TAXFOLD method, proposed by Yang et al. [23], proposes to use global and local sequential and structural features for protein fold classification. Given that an increasing number of features are proposed, simply fusing different types of feature spaces is probably not an informative means to further improve recognition accuracy. Thus, a classification method that can assess the contribution of these potentially heterogeneous object descriptors must be developed. For this reason, Damoulas et al. [21] propose a single multi-class kernel machine that informatively combines available feature groups. Apart from the SVM classifier, other single classifiers, such as RF (Random Forest) [25] and Hidden Markov Model [26], are used to construct a prediction engine in machine learning-based methods.
Chen et al. [25] recently propose an RF-based protein fold recognition method called PFP-RFSM. The framework of PFP-RFSM involves a comprehensive feature representation algorithm that can capture distinctive sequential and structural information from primary sequences and predicted structures, respectively. This feature representation algorithm generates features from seven perspectives, namely: amino acid composition, secondary structure contents, predicted relative solvent accessibility, predicted dihedral angles, PSSM matrix, nearest neighbor sequences, and sequence motifs. Features based on sequence motifs are utilized in protein fold recognition for the first time. Moreover, the PFP-RFSM method is the first to use the RF classifier as its prediction engine. As reported in [25], RF classifier is superior over the other commonly used classifiers, such as SVM, NB, and LR. In terms of overall performance, RF outperforms most of the existing methods, especially some of the ensemble-classifier methods (e.g., the well-known PFP-FunDSeqE method).
Lampros et al. [26] propose a novel optimization method for protein fold classification; the prediction model of this method is constructed based on a Markov chain trained with primary structure of proteins and on a reduced state-space HMM, which is an effective means of classifying proteins in fold categories with low computational cost. The proposed Markov chain requires only a primary sequence for training, and it is trained using a likelihood maximization algorithm. This method has proven to be effective in protein fold categorization [26].

4.2. Ensemble Classifier-Based Methods

Most of the recently developed methods for protein fold recognition are based on ensemble classifier models. Figure 3 shows the three general types of ensemble classifier models. For given n different single basic classifiers, the first type of ensemble classifier-based methods use one specific feature descriptor to encode query proteins with feature representations (Figure 3a); the feature representations are trained with each single basic classifier to create n single classifier models, and then all of the n trained single classifier models are combined with ensemble strategies to generate an ensemble classifier-based model. For a given n different single basic classifiers and n different feature descriptors, the second type of ensemble classifier-based methods use n feature descriptors to encode query proteins with n different feature representations (Figure 3b); the n feature representations are sequentially combined as one to train the n single basic classifiers, and then all of the n trained single classifier models are combined with ensemble strategies to generate an ensemble classifier-based model. For a given specific classifier and n different feature descriptors, the third type of ensemble classifier-based methods use n feature descriptors to encode query proteins with n different feature representations (Figure 3c); the n feature representations are respectively trained with a specific single classifier to construct n single classifier-based models, and then all of the n trained single classifier models are combined with ensemble strategies to generate an ensemble classifier-based model. This section highlights some representative ensemble methods used in protein fold recognition.
One well-known ensemble classifier method is PFP-FunDSeqE proposed by Shen and Chou [27]. In PFP-FunDSeqE, a novel feature extraction approach is proposed to explore the functional domain information and sequential evolution information. This approach generates 17,402 FunD features and 220 Pseudo PSSM features. The two feature groups are separately fed into an optimized evidence-theoretic K-nearest neighbor (OET-KNN) classifier to build prediction models. Accordingly, the two optimized OET-KNN models are fused to generate an ensemble classifier prediction model.
Wei et al. [13] also develop an ensemble method called PFPA. In the PFPA method, the authors design a novel feature representation algorithm considering the sequential evolutionary information and structural information. The sequential evolutionary information is derived from PSI-BLAST [28] and profiles which are generated by searching query proteins against a non-redundancy database. On the basis of the PSI-BLAST profiles, the authors compute 20 PSSM features and 420 amino acid compositional features from consensus sequences, which contain rich evolutionary information. The structural information is derived from PSI-PRED [29] profiles. To sufficiently explore the structural information, the authors calculate 27 local and 6 global secondary structure features from PSI-PRED profiles. Generally, all of the sequential and structural features are integrated to construct comprehensive feature representations of query proteins. For prediction engine construction, they build an ensemble classifier model, which fuses five basic classifier models (RF [30], NB [31], Bayes Net [32], LibSVM [33], and SMO (Sequential Minimal Optimization) [34]) with an average probability strategy. Importantly, an online webserver that implements the PFPA method is developed and freely available at http://server.malab.cn/PFPA/index.html. This method is useful to researchers in this field.
Moreover, Chen et al. [35] recently proposed a recognition method called ProFold. In ProFold, information on protein tertiary structures is first considered in its feature extraction framework in addition to other commonly used features, such as global features of amino acid sequence, PSSM features, functional domain features, and physiochemical features. The tertiary structure features are used to compute eight types of secondary structure states from PDB files by using DSSP. Additionally, ProFold proposes a novel strategy to construct an ensemble classifier. The authors first select 10 widely used basic classifiers, such as Logistic model tree [36], RF [30], LibSVM [33], Simple Logistic [36], Rotation Forest [37,38], SMO [34], NB [31], Random Tree [30], Functional tree [39], and Simple Cart [40]. Subsequently, different types of feature representations are trained using these 10 basic classifiers. For each feature type, the model with the highest accuracy is chosen, generating four single classifier models for the four feature types. The four models are DSSP model, AAsCPP model, PSSM model, and FunD model. The average probability strategy is used to fuse the four single classifier models, similar to that in the PFPA method.

5. Comparisons with Different Methods on Benchmark Dataset

To examine the effectiveness of existing machine learning-based methods in the literature for protein fold recognition, an intuitive comparison is to perform the methods on a public benchmark dataset. Here, a public and stringent dataset, proposed by Ding and Dubchak [41], is employed as a benchmark dataset for performance comparison of the existing methods. This dataset, referred as to DD, has been widely used in several studies [22,23,27,41,42,43,44,45,46,47,48,49,50]. The DD dataset is comprised of a training dataset and a testing dataset, both of which cover 27 protein fold classes in the SCOP database. The training dataset contains 311 protein sequences with ≤40% residue identity, while the testing dataset contains 383 protein sequences with ≤35% residue identity. Importantly, the sequences in the training dataset have residue identity ≤35% with that in testing dataset, thus ensuring to provide unbiased performance evaluation. The sequence distribution of each of the 27-fold classes can be seen in Table 2.
As the benchmark dataset determined, the next thing is to determine the methods for performance comparison. To provide a comprehensive comparison, we evaluated and compared the 20 representative methods published in the past 10 years (from 2006 to present) on the DD dataset. The compared 20 methods are first modeled by the training dataset of the DD dataset, and then they are tested on the testing dataset of the DD dataset. The prediction results are presented in Table 3. As shown in Table 3, we observe the following two experimental results. First, the recent ProFold exhibits the best performance among other existing methods. The overall accuracy of ProFold is 76.2%, which is 2.6%–15.7% higher than that of other methods. This demonstrates that the ProFold has great power to distinguish the 27-fold classes in the DD dataset. The significant performance improvement of ProFold contributes to the first use of the DSSP feature in the field of protein fold recognition. Their research results indicate that integrating the DSSP features into feature representations remarkably enhanced the overall accuracy from 71.2% to 76.2% [35]. This provides an alternative way to further improve predictive performance by integrating some unexplored but informative features. Second, of the 20 methods, 14 methods are based on an ensemble classifier, while 6 methods are based on a single classifier. In particular, we observe that there are 9 out of 20 methods that obtain an overall accuracy of >70%, which are PFP-FunDSeqE (70.5%), TAXFOLD (71.5%), Marfold (71.7%), Kavousi et al. (73.1%), PFPA (73.6%), Feng and Hu (70.2%), Feng et al. (70.8%), and ProFold (76.2%), respectively. Of the nine methods, only TAXFOLD is based on single classifier while the other methods are based on ensemble classifier. This indicates that ensemble classifiers are more effective than single classifiers for protein fold recognition. On the other hand, this result also explains why more recent research efforts are focused on the development of ensemble-classifier-based predictors.

6. Conclusions and Perspectives

We have systematically reviewed the recent progress in machine learning-based protein fold recognition methods. Compared with the traditional experimental methods, machine learning-based methods present three advantages. First, they demonstrate accurate, robust, and reliable performance. Second, they can be applied in large-scale protein fold recognition; this application is extremely important in the post-genomic era, wherein numerous proteins remain to be structurally characterized. Third, they can effectively address the intrinsic limitations of experimental methods, that is, their being time consuming and expensive. In the past decades, remarkable progress has been made in computational protein fold recognition. However, several challenges remain to be addressed.
First, the benchmark dataset (e.g., DD dataset) used to evaluate the performance of predictors actually suffers some limitations. For instance, the DD dataset is imbalanced. Table 2 shows that the ratio of the smallest class (“EF hand-like”) against the largest class (“immunoglobulin-like β-sandwich”) is roughly 1:4. Moreover, the sample size for each fold class is small. Only 383 training sequences belong to 27-fold classes. The largest fold class contains 30 training samples, whereas the smallest fold class contains 6 training samples. Generally, the prediction model generated based on such an imbalance and small dataset is easily overfitting.
Second, most of the existing methods, especially for those with online webservers, can only provide for the populated 27-fold class prediction. Although the sequences of the 27-fold classes cover the majority of the sequences in SCOP database, approximately 1800 protein fold classes actually exist in SCOP. Thus, developing adaptive multi-class protein fold predictors is desirable given that an increasing number of protein fold classes are being discovered.
Third, constructing informative and effective prediction engines remains a great challenge. Well-established ensemble classifiers have demonstrated their classification power in protein fold recognition. The use of deep learning algorithms for classification tasks has been a recent research hotspot in the machine learning field. Deep learning networks have been successfully applied in protein fold recognition [57]. Combining deep learning networks with well-established ensemble classifiers is probably an alternative means to improve the efficiency of protein fold recognition.
In general, machine learning-based methods can be successfully applied in protein fold recognition. In the future, machine learning methods will be extensively applied in other similar but unexplored fields, such as disease-causing amino acid change prediction [58,59,60], protein-protein binding site or interaction prediction [61,62,63], and DNA-protein binding site or interaction prediction [64,65,66].

Acknowledgments

The work was supported by the National Natural Science Foundation of China (No. 61370010).

Author Contributions

Leyi Wei participated in drafting the manuscript and performing the statistical analysis. Quan Zou participated in performing the statistical analysis. All authors read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jaroszewski, L.; Li, Z.; Cai, X.-H.; Weber, C.; Godzik, A. FFAS server: Novel features and applications. Nucleic Acids Res. 2011, 39, 38–44. [Google Scholar] [CrossRef] [PubMed]
  2. Xu, D.; Jaroszewski, L.; Li, Z.; Godzik, A. FFAS-3D: Improving fold recognition by including optimized structural features and template re-ranking. Bioinformatics 2013. [Google Scholar] [CrossRef] [PubMed]
  3. Shi, J.; Blundell, T.L.; Mizuguchi, K. Fugue: Sequence-structure homology recognition using environment-specific substitution tables and structure-dependent gap penalties. J. Mol. Biol. 2001, 310, 243–257. [Google Scholar] [CrossRef] [PubMed]
  4. Källberg, M.; Margaryan, G.; Wang, S.; Ma, J.; Xu, J. RaptorX server: A resource for template-based protein structure modeling. Protein Struct. Predict. 2014, 17–27. [Google Scholar] [CrossRef]
  5. Peng, J.; Xu, J. RaptorX: Exploiting structure information for protein alignment by statistical inference. Proteins Struct. Funct. Bioinform. 2011, 79, 161–171. [Google Scholar] [CrossRef] [PubMed]
  6. Roy, A.; Kucukural, A.; Zhang, Y. I-TASSER: A unified platform for automated protein structure and function prediction. Nat. Protoc. 2010, 5, 725–738. [Google Scholar] [CrossRef] [PubMed]
  7. Ghouzam, Y.; Postic, G.; de Brevern, A.G.; Gelly, J.-C. Improving protein fold recognition with hybrid profiles combining sequence and structure evolution. Bioinformatics 2015. [Google Scholar] [CrossRef] [PubMed]
  8. Sali, A.; Blundell, T. Comparative protein modelling by satisfaction of spatial restraints. J. Mol. Biol. 1993, 234, 779–815. [Google Scholar] [CrossRef] [PubMed]
  9. Wang, H.; He, Z.; Zhang, C.; Zhang, L.; Xu, D. Transmembrane protein alignment and fold recognition based on predicted topology. PLoS ONE 2013, 8, e69744. [Google Scholar] [CrossRef] [PubMed]
  10. Moult, J.; Fidelis, K.; Kryshtafovych, A.; Rost, B.; Hubbard, T.; Tramontano, A. Critical assessment of methods of protein structure prediction—Round VII. Proteins Struct. Funct. Bioinform. 2007, 69, 3–9. [Google Scholar] [CrossRef] [PubMed]
  11. Altschul, S.F.; Madden, T.L.; Schäffer, A.A.; Zhang, J.; Zhang, Z.; Miller, W.; Lipman, D.J. Gapped BLAST and PSI-BLAST: A new generation of protein database search programs. Nucleic Acids Res. 1997, 25, 3389–3402. [Google Scholar] [CrossRef] [PubMed]
  12. Smith, T.F.; Waterman, M.S. Identification of common molecular subsequences. J. Mol. Biol. 1981, 147, 195–197. [Google Scholar] [CrossRef]
  13. Wei, L.; Liao, M.; Gao, X.; Zou, Q. Enhanced protein fold prediction method through a novel feature extraction technique. IEEE Trans. Nanobiosci. 2015, 14, 649–659. [Google Scholar] [CrossRef] [PubMed]
  14. Bernstein, F.C.; Koetzle, T.F.; Williams, G.J.; Meyer, E.F.; Brice, M.D.; Rodgers, J.R.; Kennard, O.; Shimanouchi, T.; Tasumi, M. The protein data bank. Eur. J. Biochem. 1977, 80, 319–324. [Google Scholar] [CrossRef] [PubMed]
  15. Consortium, U. The universal protein resource (UniProt). Nucleic Acids Res. 2008, 36, D190–D195. [Google Scholar] [CrossRef] [PubMed]
  16. Kabsch, W.; Sander, C. Dictionary of protein secondary structure: Pattern recognition of hydrogen-bonded and geometrical features. Biopolymers 1983, 22, 2577–2637. [Google Scholar] [CrossRef] [PubMed]
  17. Murzin, A.G.; Brenner, S.E.; Hubbard, T.; Chothia, C. Scop: A structural classification of proteins database for the investigation of sequences and structures. J. Mol. Biol. 1995, 247, 536–540. [Google Scholar] [CrossRef]
  18. Andreeva, A.; Howorth, D.; Chothia, C.; Kulesha, E.; Murzin, A.G. SCOP2 prototype: A new approach to protein structure mining. Nucleic Acids Res. 2014, 42, 310–314. [Google Scholar] [CrossRef] [PubMed]
  19. Sillitoe, I.; Lewis, T.E.; Cuff, A.; Das, S.; Ashford, P.; Dawson, N.L.; Furnham, N.; Laskowski, R.A.; Lee, D.; Lees, J.G. Cath: Comprehensive structural and functional annotations for genome sequences. Nucleic Acids Res. 2015, 43, 376–381. [Google Scholar] [CrossRef] [PubMed]
  20. Shamim, M.T.A.; Anwaruddin, M.; Nagarajaram, H.A. Support vector machine-based classification of protein folds using the structural properties of amino acid residues and amino acid residue pairs. Bioinformatics 2007, 23, 3320–3327. [Google Scholar] [CrossRef] [PubMed]
  21. Damoulas, T.; Girolami, M.A. Probabilistic multi-class multi-kernel learning: On protein fold recognition and remote homology detection. Bioinformatics 2008, 24, 1264–1270. [Google Scholar] [CrossRef] [PubMed]
  22. Dong, Q.; Zhou, S.; Guan, J. A new taxonomy-based protein fold recognition approach based on autocross-covariance transformation. Bioinformatics 2009, 25, 2655–2662. [Google Scholar] [CrossRef] [PubMed]
  23. Yang, J.Y.; Chen, X. Improving taxonomy-based protein fold recognition by using global and local features. Proteins Struct. Funct. Bioinform. 2011, 79, 2053–2064. [Google Scholar] [CrossRef] [PubMed]
  24. Sharma, A.; Lyons, J.; Dehzangi, A.; Paliwal, K.K. A feature extraction technique using bi-gram probabilities of position specific scoring matrix for protein fold recognition. J. Theor. Biol. 2013, 320, 41–46. [Google Scholar] [CrossRef] [PubMed]
  25. Li, J.; Wu, J.; Chen, K. PFP-RFSM: Protein fold prediction by using random forests and sequence motifs. J. Biomed. Sci. Eng. 2013, 6, 1161–1170. [Google Scholar] [CrossRef]
  26. Lampros, C.; Simos, T.; Exarchos, T.P.; Exarchos, K.P.; Papaloukas, C.; Fotiadis, D.I. Assessment of optimized markov models in protein fold classification. J. Bioinform. Comput. Biol. 2014, 12, 1450016. [Google Scholar] [CrossRef] [PubMed]
  27. Shen, H.B.; Chou, K.C. Predicting protein fold pattern with functional domain and sequential evolution information. J. Theor. Biol. 2009, 256, 441–446. [Google Scholar] [CrossRef] [PubMed]
  28. Altschul, S.F.; Koonin, E.V. Iterated profile searches with PSI-BLAST—A tool for discovery in protein databases. Trends Biochem. Sci. 1998, 23, 444–447. [Google Scholar] [CrossRef]
  29. Jones, D.T. Protein secondary structure prediction based on position-specific scoring matrices. J. Mol. Biol. 1999, 292, 195–202. [Google Scholar] [CrossRef] [PubMed]
  30. Breiman, L. Random forests. Mach. Learn. 2001, 45, 5–32. [Google Scholar] [CrossRef]
  31. John, G.H.; Langley, P. Estimating continuous distributions in bayesian classifiers. In Proceedings of the Eleventh Conference on Uncertainty in Artificial Intelligence, UAI’95, Montreal, QC, Canada, 18–20 August 1995; Morgan Kaufmann Publishers Inc.: San Francisco, CA, USA, 1995; pp. 338–345. [Google Scholar]
  32. Bouckaert, R.R. Bayesian Network Classifiers in Weka; Department of Computer Science, University of Waikato: Hamilton, New Zealand, 2004. [Google Scholar]
  33. Chang, C.-C.; Lin, C.-J. LIBSVM: A library for support vector machines. ACM Trans. Intell. Syst. Technol. 2011. [Google Scholar] [CrossRef]
  34. Platt, J. Fast training of support vector machines using sequential minimal optimization. In Advances in Kernel Methods—Support Vector Learning; MIT Press: Cambridge, MA, USA, 1999; Volume 3. [Google Scholar]
  35. Chen, D.; Tian, X.; Zhou, B.; Gao, J. Profold: Protein fold classification with additional structural features and a novel ensemble classifier. BioMed Res. Int. 2016, 2016, 6802832–6802842. [Google Scholar] [CrossRef] [PubMed]
  36. Landwehr, N.; Hall, M.; Frank, E. Logistic model trees. Mach. Learn. 2005, 59, 161–205. [Google Scholar] [CrossRef]
  37. Dehzangi, A.; Phon-Amnuaisuk, S.; Manafi, M.; Safa, S. Using rotation forest for protein fold prediction problem: An empirical study. In Evolutionary Computation, Machine Learning and Data Mining in Bioinformatics, Proceedings of the 8th European Conference, EvoBIO 2010, Istanbul, Turkey, 7–9 April 2010; Springer: Berlin/Heidelberg, Germany, 2010; pp. 217–227. [Google Scholar]
  38. Rodriguez, J.J.; Kuncheva, L.I.; Alonso, C.J. Rotation forest: A new classifier ensemble method. IEEE Trans. Pattern Anal. Mach. Intell. 2006, 28, 1619–1630. [Google Scholar] [CrossRef] [PubMed]
  39. Gama, J. Functional trees. Mach. Learn. 2004, 55, 219–250. [Google Scholar] [CrossRef]
  40. Liaw, A.; Wiener, M. Classification and regression by randomforest. R News 2002, 2, 18–22. [Google Scholar]
  41. Ding, C.H.; Dubchak, I. Multi-class protein fold recognition using support vector machines and neural networks. Bioinformatics 2001, 17, 349–358. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, K.; Kurgan, L. Pfres: Protein fold classification by using evolutionary information and predicted secondary structure. Bioinformatics 2007, 23, 2843–2850. [Google Scholar] [CrossRef] [PubMed]
  43. Chen, W.; Liu, X.; Huang, Y.; Jiang, Y.; Zou, Q.; Lin, C. Improved method for predicting protein fold patterns with ensemble classifiers. Genet. Mol. Res. 2012, 11, 174–181. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, Y.; Zhang, X.; Yang, M.Q.; Yang, J.Y. Ensemble of probabilistic neural networks for protein fold recognition. In Proceedings of the 7th IEEE International Conference on Bioinformatics and Bioengineering, 2007 (BIBE 2007), Boston, MA, USA, 14–17 Ocotober 2007; IEEE: Piscataway, NJ, USA, 2007; pp. 66–70. [Google Scholar]
  45. Chmielnicki, W. A hybrid discriminative/generative approach to protein fold recognition. Neurocomputing 2012, 75, 194–198. [Google Scholar] [CrossRef]
  46. Dehzangi, A.; Phon-Amnuaisuk, S.; Dehzangi, O. Using random forest for protein fold prediction problem: An empirical study. J. Inf. Sci. Eng. 2010, 26, 1941–1956. [Google Scholar]
  47. Ghanty, P.; Pal, N.R. Prediction of protein folds: Extraction of new features, dimensionality reduction, and fusion of heterogeneous classifiers. IEEE Trans. NanoBiosci. 2009, 8, 100–110. [Google Scholar] [CrossRef] [PubMed]
  48. Lin, C.; Zou, Y.; Qin, J.; Liu, X.; Jiang, Y.; Ke, C.; Zou, Q. Hierarchical classification of protein folds using a novel ensemble classifier. PLoS ONE 2013, 8, e56499. [Google Scholar] [CrossRef] [PubMed]
  49. Nanni, L. A novel ensemble of classifiers for protein fold recognition. Neurocomputing 2006, 69, 2434–2437. [Google Scholar] [CrossRef]
  50. Shen, H.-B.; Chou, K.-C. Ensemble classifier for protein fold pattern recognition. Bioinformatics 2006, 22, 1717–1722. [Google Scholar] [CrossRef] [PubMed]
  51. Yang, T.; Kecman, V. Adaptive local hyperplane classification. Neurocomputing 2008, 71, 3001–3004. [Google Scholar] [CrossRef]
  52. Guo, X.; Gao, X. A novel hierarchical ensemble classifier for protein fold recognition. Protein Eng. Des. Sel. 2008, 21, 659–664. [Google Scholar] [CrossRef] [PubMed]
  53. Yang, T.; Kecman, V.; Cao, L.; Zhang, C.; Huang, J.Z. Margin-based ensemble classifier for protein fold recognition. Expert Syst. Appl. 2011, 38, 12348–12355. [Google Scholar] [CrossRef]
  54. Kavousi, K.; Sadeghi, M.; Moshiri, B.; Araabi, B.N.; Moosavi-Movahedi, A.A. Evidence theoretic protein fold classification based on the concept of hyperfold. Math. Biosci. 2012, 240, 148–160. [Google Scholar] [CrossRef] [PubMed]
  55. Feng, Z.; Hu, X. Recognition of 27-class protein folds by adding the interaction of segments and motif information. BioMed. Res. Int. 2014, 2014, 262850–262859. [Google Scholar] [CrossRef] [PubMed]
  56. Feng, Z.; Hu, X.; Jiang, Z.; Song, H.; Ashraf, M.A. The recognition of multi-class protein folds by adding average chemical shifts of secondary structure elements. Saudi J. Biol. Sci. 2016, 23, 189–197. [Google Scholar] [CrossRef] [PubMed]
  57. Jo, T.; Hou, J.; Eickholt, J.; Cheng, J. Improving protein fold recognition by deep learning networks. Sci. Rep. 2015, 5. [Google Scholar] [CrossRef] [PubMed]
  58. Schwarz, J.M.; Rödelsperger, C.; Schuelke, M.; Seelow, D. Mutationtaster evaluates disease-causing potential of sequence alterations. Nat. Methods 2010, 7, 575–576. [Google Scholar] [CrossRef] [PubMed]
  59. Wong, K.-C.; Zhang, Z. Snpdryad: Predicting deleterious non-synonymous human snps using only orthologous protein sequences. Bioinformatics 2014, 30, 1112–1119. [Google Scholar] [CrossRef] [PubMed]
  60. Adzhubei, I.A.; Schmidt, S.; Peshkin, L.; Ramensky, V.E.; Gerasimova, A.; Bork, P.; Kondrashov, A.S.; Sunyaev, S.R. A method and server for predicting damaging missense mutations. Nat. Methods 2010, 7, 248–249. [Google Scholar] [CrossRef] [PubMed]
  61. Guo, F.; Li, S.C.; Wang, L. Protein–protein binding sites prediction by 3D structural similarities. J. Chem. Inf. Model. 2011, 51, 3287–3294. [Google Scholar] [CrossRef] [PubMed]
  62. Guo, F.; Li, S.C.; Du, P.; Wang, L. Probabilistic models for capturing more physicochemical properties on protein–protein interface. J. Chem. Inf. Model. 2014, 54, 1798–1809. [Google Scholar] [CrossRef] [PubMed]
  63. Guo, F.; Li, S.C.; Ma, W.; Wang, L. Detecting protein conformational changes in interactions via scaling known structures. J. Comput. Biol. 2013, 20, 765–779. [Google Scholar] [CrossRef] [PubMed]
  64. Alipanahi, B.; Delong, A.; Weirauch, M.T.; Frey, B.J. Predicting the sequence specificities of DNA-and RNA-binding proteins by deep learning. Nat. Biotechnol. 2015, 33, 831–838. [Google Scholar] [CrossRef] [PubMed]
  65. Wong, K.-C.; Li, Y.; Peng, C.; Moses, A.M.; Zhang, Z. Computational learning on specificity-determining residue-nucleotide interactions. Nucleic Acids Res. 2015. [Google Scholar] [CrossRef] [PubMed]
  66. Wei, L.; Tang, J.; Zou, Q. Local-DPP: An improved DNA-binding protein prediction method by exploring local evolutionary information. Inf. Sci. 2016, in press. [Google Scholar] [CrossRef]
Figure 1. Architectures of two protein databases: SCOP and ACTH.
Figure 1. Architectures of two protein databases: SCOP and ACTH.
Ijms 17 02118 g001
Figure 2. Framework of machine learning-based methods for protein fold recognition.
Figure 2. Framework of machine learning-based methods for protein fold recognition.
Ijms 17 02118 g002
Figure 3. Construction types of ensemble-classifier models.
Figure 3. Construction types of ensemble-classifier models.
Ijms 17 02118 g003
Table 1. Summary of database sources of protein structure classification.
Table 1. Summary of database sources of protein structure classification.
Database SourcesWebsitesReferences
PDBhttp://www.rcsb.org/pdb/[14]
UniProthttp://www.uniprot.org/[15]
DSSPhttp://swift.cmbi.ru.nl/gv/dssp/[16]
SCOPhttp://scop.mrc-lmb.cam.ac.uk/[17]
SCOP2http://scop2.mrc-lmb.cam.ac.uk/[18]
CATHhttp://www.cathdb.info/[19]
Table 2. Sequence distribution of the 27-fold classes in the DD dataset.
Table 2. Sequence distribution of the 27-fold classes in the DD dataset.
IndexFold IdentifierFold NameSTrainSTestTotal
1a.1Globin-like13619
2a.3Cytochrome c7916
3a.4DNA/RNA-binding 3-helical bundle123032
4a.244-Helical up-and-down bundle7815
5a.264-Helical cytokines9918
6a.39EF hand-like6915
7b.1Immunoglobulin-like β-sandwich304474
8b.6Cupredoxin-like91221
9b.121Nucleoplasmin-like/VP161329
10b.29ConA-like lectins/glucanases7613
11b.34SH3-like barrel8816
12b.40OB-Fold131932
13b.42β-Trefoil8412
14b.47Trypsin-like serine proteases9413
15b.60Lipocalins9716
16c.1TIM β/α-barrel294877
17c.2FAD/NAD(P)-binding domain111223
18c.3Flavodoxin-like111324
19c.23NAD(P)-binding Rossmann132740
20c.37P-loop containing NTH101222
21c.47Thioredoxin-fold9817
22c.55Ribonuclease H-like motif101222
23c.69α/β-Hydrolases11718
24c.93Periplasmic binding protein-like11415
25d.15β-Grasp (ubiquitin-like)7815
26d.58Ferredoxin-like132740
27g.3Knottins (small inhibitors, toxins, lectins)132740
Total311383694
Note that STrain denotes the training dataset, and STest denotes the testing dataset.
Table 3. Performance of representative machine learning-based methods in the literature on the DD dataset.
Table 3. Performance of representative machine learning-based methods in the literature on the DD dataset.
IndexMethodsClassifier TypeReferencesOverall Accuracy (%)
1Nanni et al. (2006)Ensemble[49]61.1
2PFP-Pred (2006)Ensemble[50]62.1
3Shamim et al. (2007)Single (SVM)[20]60.5
4PFRES (2007)Ensemble[42]68.4
5Damoulas et al. (2008)Single (SVM)[21]68.1
6ALHK (2008)Ensemble[51]61.8
7GAOEC (2008)Ensemble[52]64.7
8PFP-FunDSeqE (2009)Ensemble[27]70.5
9ACCFold_AC (2009)Single (SVM)[22]65.3
10ACCFold_ACC (2009)Single (SVM)[22]66.6
11Ghanty et al. (2009)Ensemble[47]68.6
12TAXFOLD (2011)Single (SVM)[23]71.5
13Alok Sharma et al. (2012)Single (SVM)[24]69.5
14Marfold (2012)Ensemble[53]71.7
15Kavousi et al. (2012)Ensemble[54]73.1
16PFP-RFSM (2013)Single (RF)[25]73.7
17Feng and Hu (2014)Ensemble[55]70.2
18PFPA (2015)Ensemble[13]73.6
19Feng et al. (2016)Ensemble[56]70.8
20ProFold (2016)Ensemble[35]76.2

Share and Cite

MDPI and ACS Style

Wei, L.; Zou, Q. Recent Progress in Machine Learning-Based Methods for Protein Fold Recognition. Int. J. Mol. Sci. 2016, 17, 2118. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17122118

AMA Style

Wei L, Zou Q. Recent Progress in Machine Learning-Based Methods for Protein Fold Recognition. International Journal of Molecular Sciences. 2016; 17(12):2118. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17122118

Chicago/Turabian Style

Wei, Leyi, and Quan Zou. 2016. "Recent Progress in Machine Learning-Based Methods for Protein Fold Recognition" International Journal of Molecular Sciences 17, no. 12: 2118. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17122118

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop